首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
In this paper, a core–shell N-TiO2@CuOx nanomaterial with increased visible light photocatalytic activity was successfully synthesized using a simple method. By synthesizing ammonium titanyl oxalate as a precursor, N-doped TiO2 can be prepared, then the core–shell structure of N-TiO2@CuOx with a catalyst loading of Cu on its surface was prepared using a precipitation method. It was characterized in detail using XRD, TEM, BET, XPS and H2-TPR, while its photocatalytic activity was evaluated using the probe reaction of the degradation of methyl orange. We found that the core–shell N-TiO2@CuOx nanomaterial can lessen the TiO2 energy band-gap width due to the N-doping, as well as remarkably improving the photo-degradation activity due to a certain loading of Cu on the surfaces of N-TiO2 supports. Therefore, a preparation method for a novel N, Cu co-doped TiO2 photocatalyst with a core–shell structure and efficient photocatalytic performance has been provided.

In this paper, a core–shell N-TiO2@CuOx nanomaterial with increased visible light photocatalytic activity was successfully synthesized using a simple method.  相似文献   

2.
Correction for ‘Preparation and photocatalytic application of a S, Nd double doped nano-TiO2 photocatalyst’ by Shuo Wang et al., RSC Adv., 2018, 8, 36745–36753.

In the published article, Liming Bai was incorrectly not listed as the corresponding author. The correct version is shown here.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

3.
The photocatalytic reduction of CO2 is an effective way to solve the greenhouse effect. Different kinds of materials, such as semiconductors, coordination compounds, and bioenzymes, have been widely investigated to increase the efficiency of the photocatalytic reduction of CO2. However, a high selectivity and great stability are still challenges for material scientists. Here, we report for the first time visible light photocatalytic CO2 reduction by a series of CdSe/ZIF-8 nanocomposites combining the excellent CO2 adsorption capacity of ZIF-8 and the narrow energy gap of CdSe quantum dots (QDs). The composites show a higher catalytic performance than those of the pure components. Among CdSe/ZIF-8-x (x = nCdSe/nZIF-8), the highest yield (42.317 μmol g−1) for reducing CO2 to CO in 12 h, was obtained using nanocomposites with a ratio of 0.42 (nCdSe/nZIF-8) within the range of investigation.

CdSe/ZIF-8-x combines the excellent CO2 adsorption capacity of ZIF-8 and the narrow energy gap of CdSe to show an enhanced CO2 photoreduction performance.

Nowadays, global energy shortage and environmental pollution are two major obstacles to the development of human society, and have attracted increasing concern. Using solar energy to convert CO2 into valuable fuels or chemicals is extremely attractive due to its dual function of the reduction of the greenhouse effect and also as an alternative energy source to fossil fuels. Recently, different kinds of materials, such as semiconductor materials,1–3 metal complexes,4–6 and bioenzyme catalysts,7 have been explored for photocatalytic CO2 reduction.Metal–organic frameworks (MOFs) constructed from metal-containing clusters and organic building blocks are types of crystalline porous materials, and have been widely applied in many fields, such as gas storage,8 electrochemical energy storage (EES)9,10 and catalysis.11 Recently, MOFs12–14 have been considered as potential new catalysts due to their excellent capability for CO2 adsorption and capture.15 These porous materials provide a large number of catalytic active sites, and their porous structures are conducive to charge transfer.16 During the adsorption process, CO2 coordinates with unsaturated metal sites and forms chemical bonds with MOFs.12 Blom and co-workers demonstrated that CO2 can interact with metal ions and form end-on adducts with one of the oxygen lone pair orbitals.17ZIF-8, which is constructed from Zn2+ centres and imidazolate ligands, shows a high CO2 adsorption capacity since the imidazolate ligand has a high adsorption capacity for CO2 and also a strong complexation ability of CO2.18 However, ZIF-8 has a wide band gap (4.9 eV, ref. 19), which means that ZIF-8 is barely photoactive enough to catalyse CO2 reduction. However, CdSe QDs can easily be excited to generate electron–hole pairs upon visible light irradiation due to their narrow band gap. Osterloh and co-workers reported CdSe QDs of several sizes applied to photocatalytic H2 evolution and showed the quantitative relationship between the degree of quantum confinement and the photocatalytic H2 evolution.20In this work, we synthesized a series of CdSe/ZIF-8-x composites, which combine the excellent CO2 adsorption capacity of ZIF-8 with the narrow energy gap of CdSe QDs. X-ray diffraction (XRD) and energy dispersive spectroscopy (EDS) indicated the successful combination of CdSe QDs and ZIF-8. The CdSe/ZIF-8 composite exhibits an increased yield for reducing CO2 to CO compared with pure CdSe QDs or ZIF-8. Under visible light irradiation for 12 h, the CO yield was 42.317 μmol g−1, which is 6.13 and 10.84 times the yields catalysed by CdSe (6.901 μmol g−1) and by ZIF-8 (3.905 μmol g−1), respectively.Reagents used in this work were analytically pure and used without further purification. Powder X-ray diffraction (PXRD) analysis was performed using a Rigaku Dmax-2000 diffractometer equipped with a Cu Kα (λ = 0.15406 nm) radiation source. The morphology of the catalysts was observed by transmission electron microscopy (TEM, JEOL JEM-2100F) operated at 200 kV. Scanning electron microscopy (SEM) pictures were prepared using a Hitachi scanning electron microscope S-4800. Elemental mapping was carried out by energy dispersive X-ray spectroscopy (EDS) on the same instrument. Inductively coupled plasma spectrometry (ICP, Cary5000) was used for multi-elemental analyses. The CO2 absorption behaviours of the catalysts were studied with physical adsorption apparatus (ASAP 2020M). Solid UV-visible diffuse reflectance spectroscopy (UV-vis DRS) was carried out at room temperature to evaluate the band gap energy (Eg). The products of the photocatalytic CO2 reduction were detected by gas chromatography (GC7900, Techcomp). Dynamic light scattering (DLS) measurements were carried out on an Elitesizer from Brookhaven.CdSe QDs were synthesized by a previously reported procedure.21 The resulting CdSe QDs were precipitated by adding ethanol and dispersed in 5 mL of hexane as a stock solution.The synthesis of CdSe/ZIF-8 was based on the pure ZIF-8 synthesis process with modification.22 A certain quantity of the above CdSe QD stock solution was precipitated by adding ethanol, and re-dispersed in 5 mL of an n-hexanol solution of 0.1642 g (2 mmol) of 2-methylimidazole (Hmim) via ultrasonication. A solution of Zn(NO3)2·6H2O (0.074 g, 0.25 mmol) in 5 mL of n-hexanol was rapidly poured into the above solution under stirring. The product was collected by centrifugation after 1 h, washed with n-hexanol twice and dried at 80 °C for 12 h under vacuum. The samples produced from nCd2+/nZn2+ equal to 0.4, 0.8, and 1.2 were named as samples 1 to 3, respectively.The photocatalytic CO2 reduction performance of CdSe/ZIF-8-x was performed in a typical catalytic system with [Ru(bpy)3]2+ (bpy = 2′,2-bipyridine) as a photosensitizer and triethanolamine (TEOA) as a sacrificial reducing agent in CO2-saturated acetonitrile (MeCN).23–27 The photosensitizer [Ru(bpy)3]Cl2·6H2O (2 mg) and catalysts CdSe/ZIF-8-x (5 mg) were dispersed in a solution of 1 mL of triethanolamine (TEOA) and 4 mL of acetonitrile. Before irradiation, the suspension was purged with CO2 for 15 min to eliminate any air. With vigorous stirring, a 300 W Xe lamp with a 420 nm cut-off filter was utilized as the light source. After illumination for 12 h, the produced gases were analysed and quantified by gas chromatography.The molar ratios of CdSe to ZIF-8 of the composites were characterized by ICP (
SampleReaction systemProducts
10.40.30
20.80.42
31.20.59
Open in a separate windowPXRD patterns of the as-prepared samples are shown in Fig. 1. All of diffraction peaks can be indexed as CdSe with a cubic phase (PDF#19-0191, shown by *) and ZIF-8 (CCDC no. 602542). The diffraction peaks of samples 1–3 are obviously wider than those of the bulk ZIF-8 and CdSe. The diameter of CdSe is around 5 nm, which was calculated from the half-width of the diffraction peaks using Scherrer''s formula. This was also confirmed from the TEM images (Fig. 2).28Open in a separate windowFig. 1PXRD patterns of CdSe/ZIF-8-x. The peaks shown by * are the (hkl) of CdSe.Open in a separate windowFig. 2(a–c) TEM images of samples 1–3 and (d) a HRTEM image of sample 2.TEM images of samples 1–3 are shown in Fig. 2. In Fig. 2(d), the particle sizes of the CdSe QDs in the yellow circle are about 4 nm, which are the same as those of pure CdSe QDs synthesized by the same method.29 The high-resolution TEM (HRTEM) image of sample 2 (Fig. 2(d)) shows clear fringes with a lattice spacing of ca. 0.348 nm, which are attributed to the (111) plane of CdSe. This result indicates that the morphology of CdSe QDs did not change after being added to the reaction system of ZIF-8. The increasing contents of CdSe in samples 1–3 are clearly shown by the density of the QDs (Fig. 2(a–c)), which correspond with the ICP results. However, an excess addition of CdSe resulted in aggregation (Fig. 2(c)).The morphology and size of CdSe/ZIF-8-x were studied, and the SEM images are shown in Fig. S2. Clearly, the particle size of sample 2 was the smallest of the CdSe/ZIF-8-x samples. Sample 3 shows an obvious aggregation of CdSe/ZIF-8-x and an amorphous structure. The particle size was further investigated by DLS. The mean particle size values of ZIF-8 and samples 1–3 were 2443, 1279, 463, and 873 nm, respectively. Clearly, it can be seen that the moderate addition of CdSe is beneficial for smaller ZIF-8 crystals, while an excess addition of CdSe results in the aggregation of CdSe/ZIF-8-x. This result is in somewhat agreement with the TEM (Fig. 2) and SEM (Fig. S2) results.A typical EDS spectrum and elemental analysis of sample 2 are shown in Fig. S1 and Table S1, respectively, confirming the presence of Cd, Se, Zn, C and O. The elemental ratio of nCd2+/nZn2+ calculated by the EDS is only 0.09, which is lower than that of the ICP result. This is probably due to the fact that the analysis of EDS comes from the surface elements and the lower elemental ratio indicates that CdSe is wrapped inside ZIF-8. Fig. 3 shows the XPS survey spectrum and high-resolution spectra for Cd2+ 3d, Zn2+ 2p, and Se2− 3d. As shown in Fig. 3(b), the 2p3/2 and 2p1/2 binding energies of Zn2+ are located at values of 1044.8 and 1021.7 eV, respectively. Fig. 3(c) shows the 3d peak of Se2− at 54.1 eV. In addition, Fig. 3(d) shows that only two peaks appear, at binding energies of 411.7 and 404.9 eV, which are shifted towards the lower binding energy by about 0.3 eV of those of Cd2+ (3d5/2) and Cd2+ (3d3/2), from data reported in the literature.30 The above results confirm the strong combination of CdSe and ZIF-8.Open in a separate windowFig. 3(a) Survey spectrum of sample 2, (b) core level spectrum of Zn 2p, (c) core level spectrum of Se 3d, and (d) core level spectrum Cd 3d.As shown in Fig. 4, among samples 1–3, sample 2 exhibits the highest CO2 uptake at 298 K, which is about 13 times that of ZIF-8. This result means that sample 2 can greatly absorb CO2 before the reduction reaction so that it can accelerate the kinetic process of CO2 reduction. Additionally, all the samples show a linear relationship between CO2 uptake and relative pressure (0.1–1.0), indicating that the interaction between CO2 and the samples is obviously physical.27,31 According to the DLS result, the greater adsorption performance of sample 2 could be due to the relatively uniform dispersion of CdSe in ZIF-8. While the aggregation of CdSe/ZIF-8-x results in a lower CO2 uptake by the lower valid surface area and active sites from the unsaturated metal sites.12 In addition, the morphology of the samples characterized by TEM, as shown in Fig. 2, indicates that the greater adsorption performance of sample 2 is due to the sufficient quantity of CdSe and the relatively uniform dispersion of CdSe in ZIF-8 and a lower aggregation of CdSe/ZIF-8-x.Open in a separate windowFig. 4CO2 adsorption behaviour of CdSe/ZIF-8-x at 298 K.Solid UV-visible diffuse reflectance spectroscopy (UV-vis DRS) was used to evaluate the band gap energy (Eg) of samples 1–3.32 UV-vis DRS of CdSe/ZIF-8 with different proportions were studied at room temperature. From Fig. 5(a), it can be seen that the absorption wavelength of ZIF-8 is about 302 nm, which is not in the visible light region. However, the visible light absorption ability of CdSe/ZIF-8 is obviously better than that of pure ZIF-8, and the spectral response range widened to 527–630 nm. In addition, we found that the Eg value of ZIF-8 is 4.88 eV, Fig. 5(b1), which is too large to be used for visible light catalysis. However, the CdSe/ZIF-8 composites have much smaller Eg values than that of ZIF-8, at around 2.0 eV (Fig. 5(b2–b4)). This result supports the conclusion that the CdSe/ZIF-8 composites show better photocatalytic ability than ZIF-8.Open in a separate windowFig. 5(a) UV-vis DRS and (b1–b4) (Ahν)2vs. hν curves of CdSe/ZIF-8-xs.To study the thermal stability of CdSe/ZIF-8-x, we conducted TGA. As shown in Fig. S3, the initial decomposition temperatures of ZIF-8 and samples 1–3 are 118, 186, 422, and 158 °C, respectively, indicating that sample 2 has the highest thermal stability. Combined with the TEM results, the relatively uniform dispersion in sample 2 achieved the strongest combination force between CdSe and ZIF-8 among all the samples. Photocatalytic CO2 reduction experiments were carried out under visible light irradiation, and the results are summarized in ).Summary of CO2 adsorption properties and photocatalytic activities of CdSe/ZIF-8-x (with CO2 as the gas feedstock)
SampleCO2 uptake (cm3 g−1 STP)CO production rate (μmol g−1 h−1)CH4 production rate (μmol g−1 h−1)
ZIF-84.8050.3250.000
CdSe0.5750.000
Sample 14.2191.5210.092
Sample 261.4353.5260.102
Sample 34.6552.0380.056
Open in a separate windowIn a photocatalytic process, CO2 adsorption is the rate-limiting step,33 which is attributed to the fact that the CO2 conversion efficiency of the photocatalyst significantly relies on the amount of CO2 molecules adsorbed.16 Combined with the DLS result, the higher CO2 uptake of sample 2 is the main reason for the high photocatalytic activities for CO2 reduction. Theoretically, increasing the yield of CH4 is more difficult than that of CO, because reducing CO2 to CO consumes two electrons while eight electrons are needed in the CH4 transformation. Given the experimental results that CdSe/ZIF-8-x showed a higher CH4 production rate than pure CdSe and ZIF-8, we considered that ZIF-8 pores could play the role of a “nanoreactor” to enclose CO2 and CO, so as to finish the lengthy transformation process and improve the yield of CH4.34 In addition, the above inference could be supported by the CO2 adsorption capacities data (Fig. 6). As shown in Fig. 6, CdSe QDs play the core role in the photocatalytic process, as they were excited to generate electron–hole pairs upon visible light irradiation due to their narrow band gap. Furthermore, the addition of CdSe improved the conductivity of CdSe/ZIF-8-x, which attributed to the charge transfer since a good conductivity leads to only a small charge transfer resistance.35,36 ZIF-8 plays a key role as the “electrons transporter” and also as the “nanoreactor”, which means that photogenerated electrons could be transferred quickly from CdSe to ZIF-8. Then, the molecular [Ru(bpy)3]2+ photosensitizer can effectively receive the photoinduced electrons to reduce the CO2 molecule absorbed by the ZIF-8 pores to yield CO. On the other hand, the photogenerated holes are quenched by TEOA acting as a sacrificial electron donor.Open in a separate windowFig. 6Schematic illustration of the proposed mechanism of photocatalytic CO2 reduction over CdSe/ZIF-8-x.  相似文献   

4.
Correction: Expedient synthesis of eumelanin-inspired 5,6-dihydroxyindole-2-carboxylate ethyl ester derivatives     
Andrew H. Aebly  Jeffrey N. Levy  Benjamin J. Steger  Jonathan C. Quirke  Jason M. Belitsky 《RSC advances》2019,9(48):27754
Correction for ‘Expedient synthesis of eumelanin-inspired 5,6-dihydroxyindole-2-carboxylate ethyl ester derivatives’ by Andrew H. Aebly et al., RSC Adv., 2018, 8, 28323–28328.

The authors regret that an incorrect grant number was shown in the acknowledgements section of the published article. The corrected section should read:We thank the National Science Foundation (RUI grant CHE-1305919) and Oberlin College for financial support.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

5.
Correction: Nano zero valent iron (nZVI) particles for the removal of heavy metals (Cd2+, Cu2+ and Pb2+) from aqueous solutions     
Mekonnen Maschal Tarekegn  Andualem Mekonnen Hiruy  Ahmed Hussen Dekebo 《RSC advances》2021,11(43):27084
Correction for ‘Nano zero valent iron (nZVI) particles for the removal of heavy metals (Cd2+, Cu2+ and Pb2+) from aqueous solutions’ by Mekonnen Maschal Tarekegn et al., RSC Adv., 2021, 11, 18539–18551. DOI: 10.1039/D1RA01427G.

The authors regret that reference 48 was incorrect. The correct reference is given below as reference 1.1The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

6.
Correction: The selective hydrosilylation of norbornadiene-2,5 by monohydrosiloxanes     
Marina A. Guseva  Dmitry A. Alentiev  Evgeniya V. Bermesheva  Ilya A. Zamilatskov  Maxim V. Bermeshev 《RSC advances》2019,9(60):35145
  相似文献   

7.
One-pot synthesis of Cu2O/C@H-TiO2 nanocomposites with enhanced visible-light photocatalytic activity     
Aoqun Jian  Meiling Wang  Leiyang Wang  Bo Zhang  Shengbo Sang  Xuming Zhang 《RSC advances》2019,9(71):41540
As an environment-friendly semiconductor, titanium dioxide (TiO2), which can effectively convert solar energy to chemical energy, is a crucial material in solar energy conversion research. However, it has several technical limitations for environment protection and energy industries, such as low photocatalytic efficiency and a narrow spectrum response. In this study, a unique mesoporous Cu2O/C@H-TiO2 nanocomposite is proposed to solve these issues. Polystyrene beads ((C8H8)n, PS) are utilized as templates to prepare TiO2 hollow microspheres. Cu2O nanocomposites and amorphous carbon are deposited by a one-pot method on the surface of TiO2 hollow spheres. After the heterojunction is formed between the two semiconductor materials, the difference in energy levels can effectively separate the photogenerated e–h+ pairs, thereby greatly improving the photocatalytic efficiency. Furthermore, due to the visible band absorption of Cu2O, the absorption range of the prepared nanocomposites is expanded to the whole solar spectrum. Amorphous carbon, as a Cu2O reduction reaction concomitant product, can further improve the electron conduction characteristics between Cu2O and TiO2. The structure and chemical composition of the obtained nanocomposites are characterized by a series of techniques (such as SEM, EDS, TEM, XRD, FTIR, XPS, DRS, PL, MS etc.). The experimental results of the degradation of methylene blue (MB) aqueous solution demonstrate that the degradation efficiency of Cu2O/C@H-TiO2 nanocomposites is about 3 times as fast as that of pure TiO2 hollow microspheres, and a more absolute degradation can be achieved. Herein, a recyclable photocatalyst with high degradation efficiency and a whole solar spectrum response is proposed and fabricated, and would find useful applications in environment protection, and optoelectronic devices.

As an environment-friendly semiconductor, titanium dioxide (TiO2), which can effectively convert solar energy to chemical energy, is a crucial material in solar energy conversion research.  相似文献   

8.
A facile chemical synthesis of nanoflake NiS2 layers and their photocatalytic activity     
Mohammed M. Gomaa  Mohamed H. Sayed  Mahmoud S. Abdel-Wahed  Mostafa Boshta 《RSC advances》2022,12(17):10401
A single-phase and crystalline NiS2 nanoflake layer was produced by a facile and novel approach consisting of a two-step growth process. First, a Ni(OH)2 layer was synthesized by a chemical bath deposition approach using a nickel precursor and ammonia as the starting solution. In a second step, the obtained Ni(OH)2 layer was transformed into a NiS2 layer by a sulfurization process at 450 °C for 1 h. The XRD analysis showed a single-phase NiS2 layer with no additional peaks related to any secondary phases. Raman and X-ray photoelectron spectroscopy further confirmed the formation of a single-phase NiS2 layer. SEM revealed that the NiS2 layer consisted of overlapping nanoflakes. The optical bandgap of the NiS2 layer was evaluated with the Kubelka–Munk function from the diffuse reflectance spectrum (DRS) and was estimated to be around 1.19 eV, making NiS2 suitable for the photodegradation of organic pollutants under solar light. The NiS2 nanoflake layer showed photocatalytic activity for the degradation of phenol under solar irradiation at natural pH 6. The NiS2 nanoflake layer exhibited good solar light photocatalytic activity in the photodegradation of phenol as a model organic pollutant.

A single phase nanoflake NiS2 layer synthesized by a facile chemical bath deposition showed good solar light photocatalytic degradation of phenol with good stability and reusability.  相似文献   

9.
Correction: Quantum chemical elucidation of the turn-on luminescence mechanism in two new Schiff bases as selective chemosensors of Zn2+: synthesis,theory and bioimaging applications     
Jessica C. Berrones-Reyes  Blanca M. Muoz-Flores  Arelly M. Cantn-Diz  Manuel A. Treto-Surez  Dayan Pez-Hernndez  Eduardo Schott  Ximena Zarate  Víctor M. Jimnez-Prez 《RSC advances》2019,9(61):35565
  相似文献   

10.
Correction: Influence of Cu doping on the visible-light-induced photocatalytic activity of InVO4     
Natda Wetchakun  Pimonrat Wanwaen  Sukon Phanichphant  Khatcharin Wetchakun 《RSC advances》2020,10(62):37766
Correction for ‘Influence of Cu doping on the visible-light-induced photocatalytic activity of InVO4’ by Natda Wetchakun et al., RSC Adv., 2017, 7, 13911–13918, DOI: 10.1039/C6RA27138C.

The authors regret errors in Fig. 4, ,7,7, and 9 in the previously published article. The corrections for the errors in the article are described as follows:Open in a separate windowFig. 4Kubelka–Munk absorbance spectra and band gaps (insets) of the pure InVO4 (a) and 1.0 mol% Cu-doped InVO4 (b) samples.Open in a separate windowFig. 7Schematic of the charge migration and separation on Cu-doped InVO4.(1) The diffuse reflectance spectra of pure InVO4 and 1.0 mol% Cu-doped InVO4 are shown in Fig. 4. The absorption margin of 1.0 mol% Cu-doped InVO4 was shifted to a longer wavelength, indicating a decrease in the band gap with respect to pure InVO4. The absorption margins of the pure InVO4 and 1.0 mol% Cu-doped InVO4 samples were 505 nm and 510 nm, corresponding to band gaps of 2.51 eV and 2.45 eV, respectively (Fig. 4a and b).(2) The band edge positions of the conduction band (CB) and the valence band (VB) of InVO4 can be calculated by the following equation: E0CB = χEC − 0.5Eg,1 where χ is the electronegativity of the semiconductor, EC is the energy of free electrons on the hydrogen scale of 4.5 eV, Eg is the band gap of InVO4, and the χ value of InVO4 is 5.74 eV.2 The Eg value of InVO4 evaluated from the UV-vis DRS analysis was about 2.51 eV. The valence band energy (EVB) can be calculated by the following equation:3EVB = ECB + Eg, where ECB is the conduction band energy. Based on the equation above, the calculated CB and VB edge potentials of InVO4 were −0.02 eV and 2.49 eV, respectively. Now, we are in a position to discuss the photocatalytic mechanism of Cu-doped InVO4 for MB degradation (Fig. 7). In the photocatalysis process, when the absorbed photon energy () equals or exceeds the band gap, the Cu-doped InVO4 generates electron–hole (e/h+) pairs. In that case, the generated electrons from the valence band can be transferred to the conduction band of InVO4. Since the CB edge potential of InVO4 (−0.02 eV) is higher than the standard redox potential, E0(O2/O2˙) = −0.33 V vs. NHE at pH 7, this suggests that the electrons in the CB of InVO4 cannot reduce O2 to the superoxide radical ion (O2˙). In addition, the VB of InVO4 (2.49 eV) is higher than the standard redox potential, E0(OH/OH˙) = 1.99 V vs. NHE at pH 7. This indicates that the photogenerated holes in the valence band of InVO4 can oxidize the hydroxyl ion (OH) or water (H2O) to form the hydroxyl radical (OH˙).(3) Due to the contradiction between the scavenging test and the proposed photocatalytic mechanism, Fig. 9 was removed from the original article.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

11.
Hydrothermal synthesis of MoS2 nanosheet loaded TiO2 nanoarrays for enhanced visible light photocatalytic applications     
Miao Zhang  Shun Wang  Ziliang Li  Chunwang Liu  Rui Miao  Gang He  Min Zhao  Jun Xue  Zhiyuan Xia  Yongqi Wang  Zhaoqi Sun  Jianguo Lv 《RSC advances》2019,9(6):3479
A molybdenum disulfide (MoS2) nanosheet-decorated titanium dioxide (TiO2) NRA heterojunction composite was fabricated successfully through a two-step hydrothermal approach. Microstructures and optical properties of specimens were characterized by field-emission scanning electron microscopy, X-ray diffractometry, X-ray photoelectron spectroscopy, and ultraviolet-visible spectroscopy. The gaps of the TiO2 nanorods have been filled with tiny MoS2 nanosheets, which can increase the surface area of MoS2/TiO2 NRA composite thin films. In addition, the photocatalytic activity of the thin films were measured and discussed in greater detail. The appropriate hydrothermal reaction temperature of MoS2 is important for the growth of perfect MoS2/TiO2 NRA composites with significantly enhanced photocatalytic performance. The photodegradation rate and k value of MoS2-220/TiO2 are 86% and 0.0105 min−1, respectively, which are much larger than those of blank TiO2. The enhanced photocatalytic performance could be attributed to the higher visible light absorption and the reduced recombination rate of photogenerated electron–hole pairs.

A molybdenum disulfide (MoS2) nanosheet-decorated titanium dioxide (TiO2) NRA heterojunction composite was fabricated successfully through a two-step hydrothermal approach.  相似文献   

12.
Facile synthesis of an urchin-like Sb2S3 nanostructure with high photocatalytic activity     
Jing Zhou  Jiangchun Chen  Mengyao Tang  Yanqun Liu  Xiaoyu Liu  Hua Wang 《RSC advances》2018,8(33):18451
Herein, an urchin-like Sb2S3 nanostructure has been synthesized without a surfactant via a wet chemical method. The crystal structure, morphology, composition and optical properties were characterized using XRD, TEM, SEM, EDS, Raman spectroscopy, and diffuse reflectance absorption spectroscopy. The factors, including the reaction time, temperature, and ratio of the raw materials, influencing the evolution of the urchin-like morphology have been discussed, and a plausible formation mechanism for the urchin-like Sb2S3 has been proposed. The urchin-like Sb2S3 micro/nanostructure exhibits high catalytic performance towards the degradation of MB under visible light irradiation. The photodegradation ratio of MB is up to 99.32% under visible light irradiation of 130 min. Our synthesis method will be extended to prepare other photocatalysts.

Urchin-like Sb2S3 has been successfully synthesized without a surfactant using a wet chemical method. The as-prepared unique nanostructure provides a high specific surface area, leading to superior catalytic performance under visible light irradiation.

In recent years, nanomaterials with controllable size and novel morphologies have been extensively studied due to their unique physical and chemical properties.1–5 For instance, AgBr nanoplates,6 AgCl octahedrons7 and NaTi2(PO4)3 nanocubes8 show unique photocatalytic activities and electrochemical performances. Hierarchical flower-like SnO2 nanospheres9 or TiO2 nanorod arrays10,43 exhibit highly efficient gas sensing and photoelectrochemical properties.As an important V–VI compound, antimony sulfide (Sb2S3) is considered as a promising material for energy conversion due to its suitable band gap (1.5–2.2 eV), which covers the range of the solar spectrum.11–13 To the best of our knowledge, Sb2S3 nanomaterials with different morphologies have been mainly synthesized using various surfactants.14 Surfactants can regulate the morphology and structure of the nanoparticles effectively by parceling on the surface of the particles through the coordination or charge effect.15 By adding surfactants such as cetyltrimethylammonium bromide (CTAB),16 polyethylene glycol (PEG),17 dodecyltrimethylammonium bromide (DTAB)18 and polyvinylpyrrolidone (PVP),19–21 some nanostructures including straw-tied-like, nanorod22,23, nanosheet,24 nanotube, dandelion-like, double cauliflower-like,25 bar,26 dumbbell-like, and nanowire bundle structures have been successfully prepared, which show potential applications in photocatalysis and energy storage areas. However, how to effectively synthesize Sb2S3 crystals and control their morphology via a simple method is still challenging for materials scientists. Herein, we report a simple wet chemical method to synthesize an urchin-like Sb2S3 nanostructure without any surfactant. Moreover, we have examined the photocatalytic activity of Sb2S3 by the degradation of MB under simulated visible light irradiation. Our study presents an Sb2S3 nanostructure with good application in the visible-light photocatalytic field.The morphology and size of the products were revealed by the SEM and TEM images. Fig. 1a shows the SEM image of the urchins-like Sb2S3 prepared under refluxing conditions at 160 °C for 5 h. The as-prepared Sb2S3 products mainly consist of the uniform urchin-like structure. The magnified SEM image shows that the urchin-like structure consists of nanorods stretching radially from the same central point (Fig. 1b). The TEM image further confirms that the urchin-like Sb2S3 is made up of many individual nanorods with lengths about 10 μm and diameters of 100 nm (Fig. 1b and c). The measured spacing of the crystallographic planes is 0.309 nm, which is consistent with the (320) plane lattice distance of orthorhombic Sb2S3 (Fig. 1d).Open in a separate windowFig. 1(a) SEM, (b and c) TEM and (d) HRTEM images, (e) XRD pattern and (f) EDS spectrum of urchin-like Sb2S3.The XRD pattern of the obtained sample is shown in Fig. 1e. All the diffraction peaks in the XRD pattern can be readily indexed to orthorhombic Sb2S3 (JCPDS no. 42-1393) with the calculated lattice parameters of a = 1.120 nm, b = 1.128 nm and c = 0.383 nm. No characteristic peaks for impurities are observed. The fact that the reflection peaks of Sb2S3 are strong and sharp indicates that Sb2S3 is highly crystalline (Fig. 1e). The EDS spectrum reveals that the as-prepared products consist of S and Sb elements, and the observed C and Cu peaks are due to the carbon-coated Cu grid27 (Fig. 1f). Furthermore, the elemental distribution ratio of S and Sb in the compound was found to be about 3 : 2 through the quantification calculation of the EDS peaks, which was consistent with the chemical formula of antimony sulfide.28The effects of the reaction time, reaction temperature and the ratio of the raw materials on the as-obtained products were carefully studied to investigate the formation of the urchin-like morphology. Fig. 2 reveals the SEM images of the as-obtained samples prepared at different reaction times. Initially, no product was formed. When the reaction time was prolonged to 2 h, amorphous antimony sulfide was formed (Fig. 2a). After 3 h, single units with slight splitting were observed (Fig. 2b). Upon increasing the reaction time, uniform urchin-like Sb2S3 was obtained. When the reaction time was further increased to 6 h, the urchin-like structure was transformed into a rod-flower structure in which the thorns of the urchin-like structure were thicker (Fig. 2c). As shown in Fig. 2d and e, after 9 h, the rod-flower structure began to rupture and finally converted into irregular nanorods with lengths of about 15 μm, as shown in Fig. 2f.Open in a separate windowFig. 2SEM images of the Sb2S3 prepared at 160 °C for different times: (a) 2 h, (b) 3 h, (c) 6 h, (d) 9 h, (e) 12 h, and (f) 15 h.The reaction temperature also has a significant effect on the morphologies of the final products. As shown in Fig. 3a, amorphous nanoparticles with an irregular shape were formed at 140 °C. When the temperature was increased to 150 °C, a simple splitting phenomenon appeared. Upon further increasing the temperature, the urchin-like micro/nanostructure was formed. A further increase in the reaction temperature led to the formation of rod flowers. When the reaction temperature was 180 °C, some rod-like structures appeared.Open in a separate windowFig. 3SEM images of Sb2S3 prepared for 5 h at different reaction temperatures: (a) 140 °C, (b) 150 °C, (c) 170 °C and (d) 180 °C.We examined the effect of the ratio of the raw materials on the product. Only rod particles were observed at a ratio of S/Sb = 1 : 1 (Fig. 4a). When the ratio was increased to 1 : 3, the urchin shapes were observed. A further increase in the concentration of the sulfur source led to the appearance of nest structures. Irregular spherical particles were finally generated when the ratio of the raw materials was 1 : 7.Open in a separate windowFig. 4SEM images of Sb2S3 prepared with different ratios of S and Sb: (a) S/Sb = 1 : 1, (b) S/Sb = 1 : 3, (c) S/Sb = 1 : 5, and (d) S/Sb = 1 : 7.Based on the experimental results, a plausible reaction for the synthesis of the Sb2S3 crystals can be interpreted as follows:CH3CSNH2 + 2OH → S2− + CH3COO + NH4+1[Sb2(C4H4O6)2]2− → 2Sb3+ + 2(C4H4O6)4−22Sb3+ + 3S2− → Sb2S3↓3According to the abovementioned experimental results, we speculated the formation mechanism of the urchin-like Sb2S3 structure. Sb2S3 is a highly anisotropic semiconducting material with infinite ribbon-like (Sb4S6) polymers and layers in its orthorhombic crystal structure.29,30 As already reported in the literature, Sb2S3 tends to easily form one-dimensional nanorods along the c-axis due to intermolecular attraction between the antimony and sulfur atoms.31–34 Because of the fast and anisotropic crystal growth, many newborn clusters agglomerate together to deposit on a particular particle face, causing crystal splitting when the solution is in the excess saturation state.35,36 Then, each of the nanorods starts to grow into urchin-like crystals. This is similar to the chain-like crystalline structure of Bi2S3; this further ascertains that Sb2S3 has a strong splitting ability.37,38 Our research suggests that the degree of splitting is increased when the reaction time is prolonged. In the subsequent stage, the nanorods are detached from the urchin-like structures; this results in well-dispersed nanorods due to the instability of the surface energy.39 The whole process of this morphological evolution is shown schematically in Fig. 5. The amorphous nanoparticles of Sb2S3 appeared with no splitting. Upon increasing the reaction time, the Sb2S3 nanostructures change from small sheaves with simple splitting to urchin-like structures.Open in a separate windowFig. 5A schematic of the formation of the urchin-like Sb2S3 structure.The fact that the optical properties of the as-prepared Sb2S3 are determined using the absorption spectrum (Fig. 6a) provides a simple and effective method to estimate the band gap of the urchin-like Sb2S3.40,41 The band gap value of 1.64 eV determined by other researchers is quite comparable to the values reported in the bibliographical information.42 Furthermore, it can be seen that optical absorption occurs in nearly all the visible light region; this suggests that the as-synthesized Sb2S3 can be stimulated by visible light. Raman spectrum of the urchin-like Sb2S3 is shown in Fig. 6b. The appearance of sharp peaks at 147, 198, 257, 305, and 452 cm−1 suggests the formation of a highly crystalline product, which is consistent with the XRD results.43 Due to suitable band gap of urchin-like Sb2S3, the photocatalytic activity of urchin-like Sb2S3 was assessed by depositing an organic dye (MB) in an aqueous solution under visible light irradiation. The concentration of MB dye was monitored at 665 nm using an UV/vis spectrophotometer. From the blank experiment (Fig. S2), we can see a slight drop in the curve due to self-degradation of MB dye under visible light irradiation. Therefore, we can conclude that the dye is almost photo-stable under these conditions. It can be clearly observed from Fig. 6c that the organic dye is slightly degraded by Sb2S3 in the dark. According to Fig. 6d, the organic dye could be degraded by only 23.1% using H2O2, whereas the degradation efficiency for MB dye was 99.34% in the presence of Sb2S3 and H2O2 after 130 min; this indicated that the degradation of the organic dye was mainly attributed to the excellent photocatalytic activity of Sb2S3. The photoluminescence spectra of the as-prepared urchin-like Sb2S3 shows a low signal, suggesting the good charge separation properties of our samples (Fig. S1a). Fig. S1b shows the IT curves of the FTO/Sb2S3 photoelectrode, of which the photocurrent intensity is 36.057 μA cm−2 at a potential of 0 V versus Ag/AgCl, and this fast photocurrent response is consistent with the PL results. Fig. 6e shows the photocatalytic degradation of an aqueous solution of MB at different concentrations by urchin-like Sb2S3. However, with an increase in the concentration of MB, the degradation efficiency is reduced. Furthermore, the cycling performance of urchin-like Sb2S3 was measured using the same photodegradation process. The photocatalyst was obtained by centrifugation and washed with deionized water after each cycle. The degradation ratio decreased slightly after reuse for 5 times; the slight decay in the photodegradation activity was partially ascribed to the inevitable loss of the photocatalyst during the washing and centrifugation process; this suggested that Sb2S3 exhibited prominent photocatalytic stability (Fig. 6f).Open in a separate windowFig. 6(a) The diffuse reflectance absorption spectrum of the as-prepared Sb2S3 sample (the inset presents the corresponding plots of (αhυ)2versus photon energy). (b) The room-temperature Raman spectrum. (c) The degradation curve for MB in the dark. (d) The degradation curves for MB using the different catalyst systems (e) the degradation curves at different MB concentrations. (f) The cycling runs for the degradation of MB.The promising photocatalytic activity of Sb2S3 can be mainly attributed to its relatively narrow band gap (1.64 eV), making it a perfect photocatalyst to efficiently absorb and utilize visible-light energy. As is known, the urchin-like nanostructure has been extensively investigated due to its potential applications in many areas such as in varistors, electronic devices, UV-absorbers, and catalysis.47,48 This kind of nanostructure has a high specific surface area, which endows it with a large contact area with the dye and superior light harvesting efficiency in the field of photocatalysis.49 Moreover, the single crystalline nanorod-assembled urchin-like nanostructure is favorable for fast electron transfer and thus facilitates electron and hole separation.46,50,51 In particular, when urchin-like Sb2S3 is irradiated by visible light of energy greater than its band gap, electron–hole pairs will be generated and partially separated. The electrons and holes react with the adsorbed surface substances, such as O2 and OH, to form the reactive species ·O2−, ·OH, which are the major oxidative species in the decomposition of organic pollutants. Then, the oxidative species degrade the organic pollutant into small molecules such as CO2 and H2O.44,45p-Benzoquinone (BQ) and isopropyl alcohol (IPA) were used as additive agents to trap the active species ·O2 and ·OH, respectively. Fig. S2 shows the degradation curves for MB after the addition of the different capture agents. It can be seen that IPA has a significant inhibitory effect on the degradation of MB, whereas the effect of BQ is less; this indicates that ·OH plays a major role in the photocatalytic process.The degradation of the organic dye was accomplished via a series of parallel and consecutive redox reactions as show in eqn (2.1)–(2.5).Sb2S3 + → Sb2S3 (h+VB + eCB)2.1O2 + eCB → ·O22.2H2O → H+ + OH2.3OH + h+VB → ·OH2.4MB + ·OH + ·O2− → CO2 + H2O + …2.5In summary, an urchin-like Sb2S3 nanostructure was successfully prepared without surfactants using a wet chemical method. The corresponding crystal splitting formation mechanism of urchin-like Sb2S3 was tentatively suggested. The narrow band gap of the urchin-like Sb2S3 was evaluated to be 1.64 eV, which was close to the best photoelectric conversion value reported to date. In addition, urchin-like Sb2S3 exhibited excellent photocatalytic performance; the degradation efficiency for the organic dye was 99.34% after exposure to visible-light irradiation for 130 min. The high photocatalytic activity of Sb2S3 was mostly due to its narrow band gap and wide absorption range of visible light. Our investigation demonstrates that Sb2S3 with urchin-like nanostructures has great potential to be applied in the degradation of organic contaminants.  相似文献   

13.
Facile synthesis of few-layer MoS2 in MgAl-LDH layers for enhanced visible-light photocatalytic activity     
Guoyuan Zheng  Caihong Wu  Jilin Wang  Shuyi Mo  Yanwu Wang  Zhengguang Zou  Bing Zhou  Fei Long 《RSC advances》2019,9(42):24280
A new photocatalyst, few-layer MoS2 grown in MgAl-LDH interlayers (MoS2/MgAl-LDH), was prepared by a facile two-step hydrothermal synthesis. The structural and photocatalytic properties of the obtained material were characterized by several techniques including powder X-ray diffraction (XRD), Raman spectroscopy, field emission scanning electron microscopy (FESEM), high-resolution transmission electron microscopy (HRTEM), Fourier transform infrared spectroscopy (FTIR), X-ray photoelectron spectroscopy (XPS), photoluminescence spectroscopy (PL) and UV-vis absorption spectroscopy. The MoS2/MgAl-LDH composite showed excellent photocatalytic performance for methyl orange (MO) degradation at low concentrations (50 mg L−1 and 100 mg L−1). Furthermore, even for a MO solution concentration as high as 200 mg L−1, this composite also presented high degradation efficiency (>84%) and mineralization efficiency (>73%) at 120 min. The results show that the MoS2/MgAl-LDH composite has great potential for application in wastewater treatment.

Few-layer MoS2 was successfully grown in MgAl-LDH layers, utilizing the “space-confining” effect. The composite completed degraded 50 mg L−1 and 100 mg L−1 methyl orange (MO) solutions in 45 min and 105 min, respectively.  相似文献   

14.
Visible light photocatalytic one pot synthesis of Z-arylvinyl halides from E-arylvinyl acids with N-halosuccinimide     
Qiong Yu  Kun Yi Yu  Cai Feng Xu  Man-Kin Wong 《RSC advances》2022,12(7):3931
An efficient visible light photocatalytic strategy to synthesize thermodynamically less stable Z-arylvinyl halides (with up to >99/1 Z/E ratio and 86% yield) was developed. The reaction combined base-mediated halodecarboxylation of E-arylvinyl acids with N-halosuccinimide and visible light Ir-photocatalyzed isomerization of E-arylvinyl halides in a one pot sequential catalytic process.

An efficient visible light photocatalytic strategy to synthesize Z-arylvinyl halides from E-arylvinyl acids using N-halosuccinimide through a sequential halodecarboxylation/photoisomerization with up to >99/1 Z/E ratio and 86% yield under mild reaction conditions.

Visible light photocatalysis has received considerable attention in recent years owing to the mild reaction conditions, green and sustainable chemistry features, and high atom-economy.1 As reported in the literature, two different activation modes are commonly used. Most of the photochemical reactions proceed through a single-electron transfer (SET) process from the excited photosensitizers to the organic substrates or reagents. The other activation mode is an energy transfer (EnT) process in which no charge separation is involved in the whole process. This EnT activation pathway mainly depends on the triplet-state energies of the photosensitizers and the organic substrates. Synthetically useful visible light-induced organic transformation reactions through the EnT process have been successfully developed in the past decades.1n,2The synthesis of multisubstituted alkenes is an important reaction because of their versatile utility as synthetic building blocks for organic synthesis and as structural elements contributing to the significant biological properties of natural products and pharmaceuticals.3 Unlike E-alkenes, the strategies for synthesis of thermodynamically less stable Z-alkenes are not readily accessible.4 Hammond and Arai developed pioneering photochemical EZ isomerisations of stilbenes and styrenes and delineated the reaction mechanisms.5 Inspired by these mechanistic studies, visible light induced EZ isomerization has attracted great interest. In 2014, Weaver and co-workers reported Ir(ppy)3 catalyzed E to Z isomerization of allylamines proceeding via an EnT mechanism.6 In 2015, the Gilmour group developed photoisomerization of activated alkenes using (−)-riboflavin as an EnT photocatalyst.7 Furthermore, Gilmour and co-workers have reported photoisomerization of styrenyl boron species and selective isomerization of β-borylacrylates.8 In 2020, the same group reported a synthetic procedure for EZ isomerization of β-borylacrylates via EnT using thioxanthone as the sensitizer eliminating the need of an aryl unit for alkene isomerization, and the inert aryl rings were replaced by a traceless BPin handle.9Arylvinyl halides are versatile synthetic intermediates for organic synthesis. In particular, transition metal-catalyzed cross couplings of vinyl halides with organometallics, such as, organo boronic and organozinc reagents, are efficient methods for synthesis of multisubstituted alkenes.10 However, synthesis of the thermodynamically less stable Z-isomer still poses a great challenge. In 2019, Yu''s group demonstrated a synthetically useful EZ photocatalytic isomerization of styrenyl halides (Scheme 1).11 On the other hand, decarboxylation of α,β-unsaturated arylvinyl acids accompanied by simultaneous replacement by halogen is a useful reaction for the synthesis of styrenyl halides.12 Considering the importance of Z-vinyl halides in the synthesis of multisubstituted alkenes, we hypothesize that it would be interesting to combine the halodecarboxylation of α,β-unsaturated arylvinyl acids and photoisomerization of E-arylvinylhalides in a one pot sequential catalytic process. Herein, we report a novel method to synthesize Z-arylvinyl halides by visible light Ir-photocatalyzed reaction of E-arylvinyl acids with N-halosuccinimide.Open in a separate windowScheme 1Photocatalytic synthesis of multisubstituted alkenes.Our group has synthesized a series of new fluorescent quinolizinium compounds from quinolines and alkyne substrates.13 Due to the high tunability and high excited state reduction potentials of the fluorescent quinolizinium compounds, we proposed that the quinolizinium compounds could be used as photocatalysts for the synthesis of Z-arylvinyl halides from α,β-unsaturated arylvinyl acids.With this idea in mind, we started the initial investigation by treatment of (2E)-3-phenyl-2-butenoic acid (1a) with 2 equiv. of N-bromosuccinimide (NBS) in the presence of 5 mol% of photocatalyst 3a in CH3CN at room temperature under 30 W blue LEDs irradiation. 36% NMR yield of the desired product (Z/E)-(1-bromoprop-1-en-2-yl)benzene (Z/E-2a) was obtained with modest selectivity (Z/E = 44/56). Next, we optimized the reaction conditions by varying the additives. Adding K2CO3 and TBAI, E-2a was obtained only (entries 2–3, EntryPhotocatalyst (PC)AdditiveYieldb Z/Ec13a—4544/5623a1 eq. K2CO3880/10033a1 eq. TBAI750/10043a1 eq. CH3CO2H4356/4453a2 eq. CH3CO2H3577/2363a3 eq. CH3CO2H5866/3473a2 eq. PhCO2H2199/183a0.3 eq. PhCO2H3294/693b0.3 eq. PhCO2H2986/14103c0.3 eq. PhCO2H2291/9113d0.3 eq. PhCO2H3884/16123e0.3 eq. PhCO2H2488/12133f0.3 eq. PhCO2H3394/6143g0.3 eq. PhCO2H3586/14153h0.3 eq. PhCO2H4969/31163i0.3 eq. PhCO2H6360/40 Open in a separate windowaReaction conditions: treatment of E-1a (0.2 mmol), NBS (0.4 mmol) and photocatalyst (5 mol%) in 2 mL of CH3CN under N2 and blue LEDs light for 17 hours at room temperature.bYield was determined by 1H NMR using dibromomethane as internal standard.cThe Z/E ratio was determined by 1H NMR spectroscopy.Then, we screened the reaction conditions with metal photocatalyst Ir(ppy)3. Styrenyl hailde 2a was obtained in 48% NMR yield in a Z/E ratio (52 : 48) (entry 1, EntryPhotocatalyst (PC)AdditiveYieldb Z/Ec1dIr(ppy)30.3 eq PhCO2H4852/482Ir(ppy)30.3 eq PhCO2H0—3Ir(ppy)3—0—4Ir(ppy)30.3 eq. K2CO33161/395Ir(ppy)32 eq. K2CO36097/36Ir(ppy)32.5 eq. K2CO35695/57Ir(ppy)33 eq. K2CO34996/48Ir(ppy)32 eq. Na2CO35795/59Ir(ppy)32 eq. Cs2CO35494/610Ir(ppy)32 eq. K3PO46489/1111Ir(ppy)32 eq. tBuOK3892/812Ir(ppy)32 eq. DBU4794/613Ir(ppy)32 eq. Et3N967/3314Ir(Fppy)32 eq. K2CO35689/1115Ir(diFppy)32 eq. K2CO37379/2116Mes-Acr-BF42 eq. K2CO33010/9017Ru(bpy)3(PF6)22 eq. K2CO3645/9518Ru(bpy)3Cl22 eq. K2CO3480/10019eIr(ppy)32 eq. K2CO3740/10020—2 eq. K2CO3544/96Open in a separate windowaReaction conditions: treatment of E-1a (0.2 mmol), NBS (0.4 mmol) and photocatalyst (2 mol%) in 2 mL of MeOH under N2 and blue LEDs light for 17 hours at room temperature.bYield was determined by 1H NMR using dibromomethane as internal standard.cThe Z/E ratio was determined by 1H NMR spectroscopy.dThe reaction solvent was CH3CN.eNo light irradiation.After optimizing the reaction conditions, we next sought to explore the scope of the reaction. The results summarized in Open in a separate windowaReactions were performed with 0.2 mmol substrate at room temperature in MeOH (2.0 mL) using 2 mol% fac-Ir(ppy)3 under 30 W blue LEDs irradiation. Z/E ratios were determined by 1H NMR spectroscopy. Isolated yields.b5.0 mmol scale.c5 mol% fac-Ir(ppy)3.d1 mol% fac-Ir(ppy)3.To gain mechanistic insight on this reaction, the reaction progress was monitored under the optimized reaction conditions (Scheme 2). Complete halodecarboxylation of E-1c was observed in 1 hour to give E-2c, which suggests that the isomerisation is the rate-determining step. Some product Z-2c was obtained after 2 hours. Almost complete EZ isomerization was observed over the course of 16 hours.Open in a separate windowScheme 2Reaction progress monitoring of E-1c.At last, we performed two Pd-catalyzed coupling reactions with styrenyl bromide Z-2j to further illustrate the synthetic utility of this cascade reaction (Scheme 3). When 4-methoxyphenylboronic acid was treated with bromide Z-2j using Pd(PtBu3)2 as catalyst, Suzuki–Miyaura cross coupling reaction14 successfully afforded trisubstituted alkene Z-4 (Z/E = 98/2) in 78% yield. Moreover, Sonogashira coupling reaction15 with 1-ethynyl-4-methylbenzene gave enyne Z-5 (Z/E = 98/2) in 87% yield.Open in a separate windowScheme 3Pd-catalyzed coupling reactions of styrenyl bromide Z-2j.  相似文献   

15.
Facile one-pot synthesis of Mg-doped g-C3N4 for photocatalytic reduction of CO2     
Xinyue Dong  Suicai Zhang  Hualin Wu  Zhuo Kang  Li Wang 《RSC advances》2019,9(49):28894
Graphitic carbon nitride (g-C3N4) has attracted wide attention due to its potential in solving energy and environmental issues. However, rapid charge recombination and a narrow visible light absorption region limit its performance. In our study, Mg-doped g-C3N4 was synthesized through a facile one-pot strategy for CO2 reduction. After Mg doping, the light utilization efficiency and photo-induced electron–hole pair separation efficiency of the catalysts were improved, which could be due to the narrower band gap and introduced midgap states. The highest amounts of CO and CH4 were obtained on Mg-CN-4% under ultraviolet light illumination, which were about 5.1 and 3.8 times that of pristine g-C3N4, respectively; the yield of CO and CH4 reached 12.97 and 7.62 μmol g−1 under visible light irradiation. Our work may provide new insight for designing advanced photocatalysts in energy conversion applications.

Graphitic carbon nitride (g-C3N4) has attracted wide attention due to its potential in solving energy and environmental issues.  相似文献   

16.
Vortex fluidic mediated synthesis of polysulfone     
Aghil Igder  Scott Pye  Ahmed Hussein Mohammed Al-Antaki  Alireza Keshavarz  Colin L. Raston  Ata Nosrati 《RSC advances》2020,10(25):14761
Polysulfone (PSF) was prepared under high shear in a vortex fluidic device (VFD) operating in confined mode, and its properties compared with that prepared using batch processing. This involved reacting the pre-prepared disodium salt of bisphenol A (BPA) with a 4,4′-dihalodiphenylsulfone under anhydrous conditions. Scanning electron microscopy (SEM) established that in the thin film microfluidic platform, the PSF particles are sheet-like, for short reaction times, and fibrous for long reaction times, in contrast to spherical like particles for the polymer prepared using the conventional batch synthesis. The operating parameters of the VFD (rotational speed of the glass tube, its tilt angle and temperature) were systematically varied for establishing their effect on the molecular weight (Mw), glass transition temperature (Tg) and decomposition temperature, featuring gel permeation chromatography (GPC), differential scanning calorimetry (DSC) and thermal gravimetric analysis (TGA) respectively. The optimal VFD prepared PSF was obtained at 6000 rpm rotational speed, 45° tilt angle and 160 °C, for 1 h of processing with Mw ∼10 000 g mol−1, Tg ∼158 °C and decomposition temperature ∼530 °C, which is comparable to the conventionally prepared PSF.

Polysulfone (PSF) was prepared under high shear in a vortex fluidic device (VFD) operating in confined mode. This involved reacting the pre-prepared disodium salt of bisphenol A (BPA) with a 4,4′-dihalodiphenylsulfone under anhydrous conditions.  相似文献   

17.
Correction: Synthesis of Bi2WO6/Na-bentonite composites for photocatalytic oxidation of arsenic(iii) under simulated sunlight     
Quancheng Yang  Yunxiang Dai  Zijian Huang  Jing Zhang  Ming Zeng  Changsheng Shi 《RSC advances》2019,9(70):40810
Correction for ‘Synthesis of Bi2WO6/Na-bentonite composites for photocatalytic oxidation of arsenic(iii) under simulated sunlight’ by Quancheng Yang et al., RSC Adv., 2019, 9, 29689–29698.

Ref. 28 in the published article was incorrect, with an incorrect page range provided. The correct version is shown as ref. 1 below.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

18.
Correction: Cu(ii)/Ni(ii)–organic frameworks constructed from the homometallic clusters by 5-(2-carboxyphenoxy)isophthalic acid and N-ligand: synthesis,structures and visible light-driven photocatalytic properties     
Qi-Wei Xu  Qiu-Shuang Wang  Shan-Shan Li  Xia Li 《RSC advances》2019,9(30):17414
Correction for ‘Cu(ii)/Ni(ii)–organic frameworks constructed from the homometallic clusters by 5-(2-carboxyphenoxy)isophthalic acid and N-ligand: synthesis, structures and visible light-driven photocatalytic properties’ by Qi-Wei Xu et al., RSC Adv., 2019, 9, 16305–16312.

In the Acknowledgements section, a funder, Laboratory Open Project Fund of Capital Normal University (LIP18009S047) was omitted.The revised acknowledgements should read:The authors are grateful to the Key Project of Science and Technology plan of Beijing Education Commission (KZ201910028038), the National Natural Science Foundation of China (21471104) and Laboratory Open Project Fund of Capital Normal University (LIP18009S047).The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

19.
Base mediated spirocyclization of quinazoline: one-step synthesis of spiro-isoindolinone dihydroquinazolinones     
Rapolu Venkateshwarlu  V. Narayana Murthy  Krishnaji Tadiparthi  Satish P. Nikumbh  Rajesh Jinkala  Vidavalur Siddaiah  M. V. Madhu babu  Hindupur Rama Mohan  Akula Raghunadh 《RSC advances》2020,10(16):9486
A novel approach for the spiro-isoindolinone dihydroquinazolinones has been demonstrated from 2-aminobenzamide and 2-cyanomethyl benzoate in the presence of KHMDS as a base to get moderate yields. The reaction has been screened in various bases followed by solvents and a gram scale reaction has also been executed under the given conditions. Based on the controlled experiments a plausible reaction mechanism has been proposed. Further the substrate scope of this reaction has also been studied.

A novel approach for the spiro-isoindolinone dihydroquinazolinones has been demonstrated from 2-aminobenzamide and 2-cyanomethyl benzoate in the presence of KHMDS as a base to get moderate yields.

Because of their strong potent biologically activity, heterocyclic compounds have been a constant source of inspiration for the invention of new drugs especially for pharmaceutical and agro chemical industries.1 Indeed, investigation of novel methods for the synthesis of various natural products and heterocycles has always been a challenging task in modern organic chemistry. Amid all, spiro based scaffolds have been found to be very interesting because of their structural diversity. In spite of their intrinsic structures and immense biological activity there is a tremendous demand for the chemistry of spiro-isoindolinone dihydroquinazolinones.2 Indeed, nitrogen containing heterocyclic compounds like spiro-oxindole, spiro-isoindoline, spiro-isoindolinone are playing a significant role in medicinal chemistry and synthetic transformations.3 Moreover these compounds present in many natural products as a core unit like Lennoxamine, Zopiclone, Taliscanine and Pazinaclone (Fig. 1). In addition, many unnatural spiro-isoindolinones show significant biological activities acting as anti HIV-1, antiviral, antileukemic, anesthetic and antihypertensive agents.4,5 Notably, the spiro-isoindolinone dihydroquinazolinone unit has been found to be a combination of two potent pharmacophore units of dihydroquinazolinone and spiro-isoindolinone. Inspite of their remarkable biological activity afore mentioned, various methods have been developed for their synthesis like lithiation approaches, base mediated protocols, Diels–Alder and Wittig reactions, electrophilic and radical cyclization, metal-catalysed reactions and various electrochemical procedures.6Open in a separate windowFig. 1Some biologically active spiro-isoindolinone and quinazolinone units.Previously, a number of metal catalyzed reactions have also been reported for the spiroannulations.7 Among all, Nishimura et al. developed an Ir(i) catalyzed [3 + 2] annulation of benzosultam and N-acylketimines with 1,3-dienes via C–H activation for the synthesis of aminocyclopentene derivatives. Further, Xingwei Li et al. developed a Rh(iii)-catalysed [3 + 2] annulation of cyclic N-sulfonyl or N-acyl ketimines with activated alkenes for the preparation of various spirocyclic compounds.8 Recently Yangmin Ma et al. developed a one pot nano cerium oxide catalyzed synthesis of spiro-oxindole dihydroquinazolinone derivatives (Scheme 1).5c However, development of these type of novel compounds is always challenging and more attractive. Indeed, to the best of our knowledge there are no reports for the synthesis of spiro-isoindolinone dihydroquinazolinones. This led us to give more attention to study these compounds.Open in a separate windowScheme 1Different strategies for the synthesis of quinazolinone units.In continuation of our earlier efforts9 for the synthesis of various dihydroquinazolinones, herein we would like to report KHMDS mediated synthesis of novel spiro-isoindolinone dihydroquinazolinones. We envisioned the retro synthetic pathway for these compounds, as depicted in Scheme 2. Accordingly these compounds could be synthesised from 2-aminobenzamide and methyl-2-cyanobenzoate or ethyl-2-cyanobenzoate.Open in a separate windowScheme 2Retro synthetic approach for the synthesis of quinazolinone unit.Indeed, in order to understand the reaction conditions, we have commenced the reaction by taking 2-amino-N-hexyl-benzamide (2) and methyl-2-cyanobenzoate (3) as a model substrates. However, in the initial phase of reaction optimisation, we have screened the reaction in different bases (Fig. 2) and to our delight amongst all the bases KHMDS, LiHMDS and NaHMDS were amenable to get moderate yields. However the reaction had not progressed at low temperatures (5 °C) and could improve the yield at room temperature. Moreover the reaction underwent complete conversion with 1.5 equivalents of base. Further, the reaction was also executed with ethyl-2-cyanobenzoate and could replicate the same yield. Incontinuation, the reaction was futile when the reaction was carried out in DIPEA, DBU, K2CO3 and Cs2CO3. Subsequently, the reaction in NaOMe and tBuOK produced exclusively the hydrolysis product (4) of methyl-2-cyanobenzoate (Open in a separate windowFig. 2Base screening in 1,4-dioxane.Optimisation of the base-mediated spiroannulationa
EntryBaseSolvent1jb (%)Cyano benzoic acid (4)(%)
1KHMDS1,4-Dioxane604
2NaHMDS1,4-Dioxane583
3NaOMe1,4-Dioxane60
4 t BuOK1,4-Dioxane60
5KHMDSTHF405
6KHMDS1,2-DME605
7LiHMDS1,4-Dioxane555
Open in a separate windowaReaction conditions: KHMDS (1 M, 1.5 mmol), 2-amino-benzamide (1 mmol) and methyl-2-cyanobenzoate (1.5 mmol) in 1,4-dioxane (10 mL).bIsolated yield.Gratifyingly, among all the solvents 1,4-dioxane, THF and 1,2-dimethoxyethane were found to get good to moderate yields. Whereas, other solvents like DCM ended up with non-polar spots where as in toluene unknown polar impurity was observed. However, there is no reaction progress observed in the presence of trifluoroacetic acid as well as in BF3·Et2O as a solvent (Fig. 3).Open in a separate windowFig. 3Solvent screening in the presence of KHMDS.With the optimized conditions in hand, we have explored the applicability of our reaction with various substrates by taking various groups like alkyl, cyclopropyl, cyclohexyl, cycloheptyl, benzyl, naphthyl, furan and to our delight all the substrates were well tolerated under the aforementioned optimal conditions ( Open in a separate windowaReaction conditions: KHMDS (1 M, 1.5 mmol), 2-amino-benzamide (1 mmol) and methyl/ethyl-2-cyanobenzoate (1.5 mmol) in 1,4-dioxane (10 mL).Based on the aforementioned studies and the literature reports, a plausible mechanism for this reaction has been predicted (Scheme 3). Indeed, to gain insight into the mechanism a series of control experiments have been executed under the similar reaction conditions. Initially the reaction has been carried out without base and both the starting materials were intact. Further the reaction without 2-aminobenzamide resulted hydrolysis product. To explore further, the reaction has also been executed on a 10 gram-scale for the synthesis of 1j and has successfully been demonstrated under the aforementioned optimized conditions.Open in a separate windowScheme 3Plausible mechanisms for the synthesis of spiro-isoindolinone dihydroquinazolinones.The Scheme 3 describes a plausible mechanism for the preparation of compound 1. Initially, KHMDS will abstract N–H proton of amide and nucleophilic nitrogen will attack the cyanobenzoate to get imine intermediate 6 and 7, which on subsequent cyclization lead to the formation of 8 and 9. Finally, the compounds 8 and 9 underwent cyclization to get the spiro-isoindolinone dihydroquinazolinone 1.  相似文献   

20.
Correction: Solvent-controlled synthesis of multicolor photoluminescent carbon dots for bioimaging     
Yang Yan  Longyu Xia  Lan Ma 《RSC advances》2019,9(46):26551
Correction for ‘Solvent-controlled synthesis of multicolor photoluminescent carbon dots for bioimaging’ by Yang Yan et al., RSC Adv., 2019, 9, 24057–24065.

The authors regret that the contributions of the authors were not correctly indicated in the original article. Yang Yan should be designated as the sole first author, and therefore the footnote stating “The two authors contributed equally” was in error. The correct author list is as shown above.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号